Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 25
Filter
Add more filters










Publication year range
1.
Mass Spectrom Rev ; 2023 Jan 31.
Article in English | MEDLINE | ID: mdl-36721337

ABSTRACT

The evolution and applications of flow tube mass spectrometry in the study of catalysis promoted by atomic metal ions are tracked from the pioneering days in Boulder, Colorado, to the construction and application of the ICP/SIFT/QqQ and ESI/qQ/SIFT/QqQ instruments at York University and the VISTA-SIFT instrument at the Air Force Research Laboratory. The physical separation of various sources of atomic metal ions from the flow tube in the latter instruments facilitates the spatial resolution of redox reactions and allows the separate measurement of the kinetics of both legs of a two-step catalytic cycle, while also allowing a view of the catalytic cycle in progress downstream in the reaction region of the flow tube. We focus on measurements on O-atom transfer and bond activation catalysis as first identified in Boulder and emphasize fundamental aspects such as the thermodynamic window of opportunity for catalysis, catalytic efficiency, and computed energy landscapes for atomic metal cation catalysis. Gas-phase applications include: the catalytic oxidation of CO to CO2 , of H2 to H2 O, and of C2 H4 to CH3 CHO all with N2 O as the source of oxygen; the catalytic oxidation of CH4 to CH3 OH with O3 ; the catalytic oxidation of C6 H6 with O2 . We also address the environmentally important catalytic reduction of NO2 and NO to N2 with CO and H2 by catalytic coupling of two-step catalytic cycles in a multistep cycle. Overall, the power of atomic metal cations in catalysis, and the use of flow tube mass spectrometry in revealing this power, is clearly demonstrated.

2.
J Am Soc Mass Spectrom ; 33(8): 1419-1426, 2022 Aug 03.
Article in English | MEDLINE | ID: mdl-35533366

ABSTRACT

The interaction of the atomic coinage metal cations Cu+, Ag+, and Au+ with O2, a weak ligand, and C6H6, a strong ligand, was investigated with measurements of rate coefficients of ligation and quantum-chemical computations of ligation energies with an eye on relativistic effects going down the periodic table. Strong "third row enhancements" were observed for both the rate coefficients of ligation and ligation energies with the O2 ligand and for the formation of both the mono- and bis-adducts of M+ and the monoadduct of M+(C6H6). The computations revealed that the third-row enhancement in the ligation energy is attributable to a relativistic increase in the ligation energy. This means that rate coefficient measurements down the periodic table for the ligation of coinage metal cations with O2 provide a probe of the relativistic effect in ligation reactions, as expected from the known dependence of the rate coefficient of ligation on the ligation energy. The much stronger benzene ligand was observed to ligate the atomic coinage metal cations with nearly 100% efficiency so that there is no, or only slightly, visible third-row enhancement despite the strong relativistic effect in the binding energy that is revealed by the calculations. Relativistic effects contribute substantially to the extraordinary stability against deligation of all the observed mono- and bis-adducts of Au+ relative to Ag+, truly a "third-row enhancement".

3.
J Am Soc Mass Spectrom ; 30(10): 1850-1856, 2019 Oct.
Article in English | MEDLINE | ID: mdl-31111415

ABSTRACT

Mono-ligation kinetics were measured for ammonia reacting with atomic cations in the first two groups of the periodic table (K+, Rb+, Cs+ and Ca+, Sr+, Ba+). Also, mono-ligation energies were computed using density functional theory (DFT) in an attempt to assess the role of non-covalent electrostatic interactions in these chemical reactions. The measurements were performed at room temperature in helium bath gas at 0.35 Torr using an inductively coupled plasma/selected-ion flow tube (ICP/SIFT) tandem mass spectrometer. Rate coefficients are reported for ammonia addition, the only reaction channel that was observed with all these cations. A systematic decrease in the rate of addition of NH3 was observed for both group 1 and 2 cations going down the periodic table. The computational studies predict a decrease in the adduct binding energy and an increase in the bond separation going down groups 1 and 2 of the periodic table and provide some insight into the role of the extra selectron in the group 2 radical cations in ligand bonding. A correlation is seen between the efficiency of ligation and the binding energy of the adduct ion and attributed to the lifetime of the intermediate encounter complex against back dissociation which is dependent on its well depth. Higher-order additions of ammonia were also observed. Remarkable differences in the extent and kinetics were seen between the group 1 and 2 cations, and these were attributed to the occurrence of ammonia solvation of the extra s electron in the higher-order adducts of the alkaline earth cations.

4.
Eur J Mass Spectrom (Chichester) ; 25(1): 44-49, 2019 Feb.
Article in English | MEDLINE | ID: mdl-30773917

ABSTRACT

The kinetics of ammonia ligation to atomic first and second row transition metal cations were measured in an attempt to assess the role of ligand field effects in gas-phase ion-molecule reaction kinetics. Measurements were performed at 295 ± 2 K in helium bath gas at 0.35 Torr using an inductively coupled plasma/selected-ion flow tube tandem mass spectrometer. The atomic cations were produced at ca. 5500 K in an inductively coupled plasma source and were allowed to decay radiatively and to thermalize by collisions with argon and helium atoms prior to reaction. A strong correlation was observed across the periodic table between the measured rate coefficients for ammonia ligation and measured/calculated bond dissociation energies. A similar strong correlation is seen with the ligand field stabilization energy. So ligand field stabilization energies should provide a useful predictor of relative rates of ligation of atomic metal ions.

5.
J Am Soc Mass Spectrom ; 25(3): 490-7, 2014 Mar.
Article in English | MEDLINE | ID: mdl-24425289

ABSTRACT

The optimization of ion/molecule chemistry in a differential mobility spectrometer (DMS) is shown to result in improved peak capacity, separation, and sensitivity. We have experimented with a modifier composed of multiple components, where each component accomplishes a specific task on mixtures of peptides and small drug molecules. Use of a higher proton affinity modifier (hexanol) provides increased peak capacity and separation. Analyte ion/modifier proton transfer is suppressed by adding a large excess of low proton affinity modifier (water or methanol), significantly increasing signal intensity and sensitivity for low proton affinity analytes. Finally, addition of an electrical arcing suppressant (chloroform) allows the device to operate reliably at higher separation fields, improving peak capacity and separation. We demonstrate a 20% increase in the device peak capacity without any loss of sensitivity and estimate that further optimization of the modifier composition can increase this to 50%. Use of 3-, 4-, or even 5-component modifiers offers the opportunity for the user to fine-tune the modifier performance to maximize the device performance, something not possible with a single component modifier.


Subject(s)
Mass Spectrometry/methods , Peptide Fragments/analysis , Pharmaceutical Preparations/analysis , Animals , Cattle , Organic Chemicals , Peptide Fragments/chemistry , Pharmaceutical Preparations/chemistry , Protons , Serum Albumin, Bovine
7.
J Phys Chem A ; 117(18): 3786-90, 2013 May 09.
Article in English | MEDLINE | ID: mdl-23581853

ABSTRACT

In a recent publication by J. J. Melko et al. (J. Phys. Chem. A2012, 116, 11500-11508) on the reactions of Fe(+) cations with NO and NO2, these authors made a number of assertions regarding the work previously published in our laboratory. Melko et al. assert that our previously reported data was erroneously analyzed, resulting in our misreporting of the Fe(+) + NO2 reaction branching ratio for NO(+). Also, they proposed that this alleged misreporting made it likely for the second-order chemistry observed in our Fe(+) + NO experiments to be a product of an impurity of NO2 in our NO reagent and, furthermore, that our reported rate coefficient for the effective second-order chemistry was unreasonably high on the basis of their model calculations. Despite extensive private communications in which we presented detailed data supporting our original data analysis to Melko et al., these authors proceeded to publish their critique without any reference to this data. Here, we present the data communicated by us to Melko et al. and show that our result reported earlier for the Fe(+) + NO2 reaction branching ratio to form NO(+) is accurate and, furthermore, that there is no evidence for a sufficient NO2 impurity in any of our NO experiments. We suggest that the discrepancy in the results observed by us and Melko et al. may be attributed to a reaction with the dimer (NO)2. This possibility was dismissed in our earlier work as the dimer concentration under the flow tube conditions was calculated to be below 10(-5)% of the monomer, but the new results of J. J. Melko et al. raise the dimer reaction as a real possibility. Finally, J. J. Melko et al. appear to have misunderstood the mechanism of the second-order NO chemistry that we had proposed.

8.
J Breath Res ; 7(2): 026005, 2013 Jun.
Article in English | MEDLINE | ID: mdl-23579200

ABSTRACT

Volatile organic compounds (VOCs) in exhaled human breath can serve as potential disease-specific and exposure biomarkers and therefore can reveal information about a subject's health and environment. Pyridine, a VOC marker for exposure to tobacco smoke, and isoprene, a liver disease biomarker, were studied using atmospheric pressure chemical ionization mass spectrometry (APCI-MS). While both molecules could be detected in low-ppb levels, interactions of the ionized analytes with their neutral forms and ambient air led to unusual ion/molecule chemistry. The result was a highly dynamic system and a nonlinear response to changes in analyte concentration. Increased presence of ambient water was found to greatly enhance the detection limit of pyridine and only slightly decrease that of isoprene. APCI-MS is shown to be a promising analytical tool in breath analysis with good detection limits, but its application requires a better understanding of the ion/molecule chemistry that may affect VOC quantification from a chemically complex system such as human breath.


Subject(s)
Breath Tests/methods , Butadienes/analysis , Environmental Illness/diagnosis , Gases/chemistry , Hemiterpenes/analysis , Mass Spectrometry/methods , Pentanes/analysis , Pyridines/analysis , Atmospheric Pressure , Biomarkers/analysis , Environmental Illness/metabolism , Exhalation , Humans , Volatile Organic Compounds/analysis
9.
J Phys Chem A ; 117(6): 1151-7, 2013 Feb 14.
Article in English | MEDLINE | ID: mdl-22894764

ABSTRACT

Experimental results are reported for the gas-phase room-temperature kinetics of chemical reactions between nitrogen dioxide (NO(2)) and 46 atomic main-group and transition metal cations (M(+)). Measurements were taken with an inductively-coupled plasma/selected-ion flow tube (ICP/SIFT) tandem mass spectrometer in helium buffer gas at a pressure of 0.35 ± 0.01 Torr and at 295 ± 2 K. The atomic cations were produced at ca. 5500 K in an ICP source and allowed to decay radiatively and to thermalize to room temperature by collisions with Ar and He atoms prior to reaction with NO(2). Measured apparent bimolecular rate coefficients and primary reaction product distributions are reported for all 46 atomic metal cations and these provide an overview of trends across and down the periodic table. Three main types of reactions were observed: O-atom transfer to form either MO(+) or NO(+), electron transfer to form NO(2)(+), and addition to form MNO(2)(+). Bimolecular O-atom transfer was observed to predominate. Correlations are presented between reaction efficiency and the O-atom affinity of the metal cation and between the prevalence of NO(+) product formation and the electron recombination energy of the product metal oxide cation. Some second-order reactions are evident with metal cations that react inefficiently. Most interesting of these is the formation of the MNO(+) cation with Rh(+) and Pd(+). The higher-order chemistry with NO(2) is very diverse and includes the formation of numerous NO(2) ion clusters and a number of tri- and tetraoxide metal cations. Group 2 metal dioxide cations (CaO(2)(+), SrO(2)(+), BaO(2)(+)) exhibit a unique reaction with NO(2) to form MO(NO)(+) ions perhaps by NO transfer from NO(2) concurrent with O(2) formation by recombination of a NO(2) and an oxide oxygen.


Subject(s)
Helium/chemistry , Nitrogen Dioxide/chemistry , Temperature , Transition Elements/chemistry , Cations/chemistry , Kinetics
10.
Chemistry ; 16(38): 11605-10, 2010 Oct 11.
Article in English | MEDLINE | ID: mdl-20827690

ABSTRACT

The room-temperature chemical kinetics has been measured for the catalytic activity of Group 10 atomic cations in the oxidation of methane to methanol by ozone. Ni(+) is observed to be the most efficient catalyst. The complete catalytic cycle with Ni(+) is interpreted with a computed potential energy landscape and, in principle, has an infinite turnover number for the oxidation of methane, without poisoning side reactions. The somewhat lower catalytic activity of Pd(+) is reported for the first time and also explored with DFT calculations. Pt(+) is seen to be ineffective as a catalyst because of the observed failure of PtO(+) to convert methane to methanol.

11.
Phys Chem Chem Phys ; 12(18): 4852-62, 2010 May 14.
Article in English | MEDLINE | ID: mdl-20428568

ABSTRACT

Results of experimental investigations are reported for the gas-phase kinetics of chemical reactions between nitrogen dioxide (NO(2)) and 14 different atomic cations of the lanthanide series, Ln(+) (Ln = La-Lu, excluding Pm), and their monoxides, LnO(+). Measurements were taken with an inductively-coupled plasma/selected-ion flow tube (ICP/SIFT) tandem mass spectrometer in helium buffer-gas at a pressure of 0.35 +/- 0.01 Torr and at 295 +/- 2 K. The atomic lanthanide cations were produced at ca. 5500 K in an ICP source and allowed to decay radiatively and to thermalize by collisions with Ar and He atoms prior to reaction with NO(2). The atomic ions were observed to react rapidly with NO(2) with large rate coefficients, k > 2 x 10(-10) cm(3) molecule(-1) s(-1), and almost exclusively by oxygen-atom abstraction to produce lanthanide-oxide LnO(+) cations. In contrast to results of previous studies with many other molecules, the reaction efficiency exhibits essentially no dependence upon the energy required to promote an electron to achieve a d(1)s(1) excited electronic configuration, in which two non-f electrons are available to Ln(+) for chemical bonding. Apparently the radical character of NO(2) (X (2)A(1)) leads to the efficient formation of LnO(+) by the end-on abstraction of an oxygen atom by Ln(+). In the reactions with La(+), Ce(+), Pr(+) and Gd(+) an additional minor channel (less than 2%) leads to the formation of NO(+). The LnO(+) product ions participate in various secondary and higher order reactions with NO(2) resulting in the formation of ions of the type LnO(x)(NO)(y)(NO(2))(z)(+) with x = 1-2, y = 0-2, and z = 0-2, as well as the ions NO(+) and NO(2)(+).

12.
J Phys Chem A ; 114(1): 241-6, 2010 Jan 14.
Article in English | MEDLINE | ID: mdl-19848400

ABSTRACT

Reactions of (14) atomic lanthanide cations (excluding Pm(+)) with ammonia have been surveyed in the gas phase by using an inductively coupled plasma/selected-ion flow tube (ICP/SIFT) tandem mass spectrometer to measure rate coefficients and product distributions in He at 0.35 +/- 0.01 Torr and 295 +/- 2 K. Primary reaction channels were observed corresponding to H(2) elimination with formation of the protonated lanthanum nitride and NH(3) addition. H(2) elimination was seen only in the reactions with La(+), Ce(+), Gd(+), and Tb(+) and occurs with these ions exclusively. NH(3) addition was seen with Pr(+), Nd(+), Sm(+), Eu(+), Dy(+), Ho(+), Er(+), Tm(+), Yb(+), and Lu(+). Higher-order sequential addition of up to five NH(3) molecules was observed with the Ln(+)(NH(3)) and LnNH(+) ions. The reaction efficiency of the primary reactions is seen to decrease as the energy required to promote an electron to make two non-f electrons available for bonding increases. The periodic trend in reaction efficiency along the lanthanide series matches quite closely the periodic trend in the electron-promotion energy required to achieve a d(1)s(1) or d(2) excited electronic configuration in the lanthanide cation. With La(+), Ce(+), Gd(+), and Tb(+), the electrostatic attraction between the atomic lanthanide cation and ammonia is sufficiently strong to provide enough energy to achieve electron promotion and to overcome any barriers to subsequent N-H bond insertion and H(2) loss, but this is not the case with the other lanthanide cations with which collisional stabilization of the intermediate adduct ion, with or without insertion of Ln(+), predominates.


Subject(s)
Ammonia/chemistry , Gases/chemistry , Lanthanoid Series Elements/chemistry , Temperature , Cations/chemistry , Kinetics , Periodicity , Tandem Mass Spectrometry
13.
J Phys Chem A ; 113(19): 5602-11, 2009 May 14.
Article in English | MEDLINE | ID: mdl-19371125

ABSTRACT

Reactions of methane have been measured with 59 atomic metal cations at room temperature in helium bath gas at 0.35 Torr using an inductively-coupled plasma/selected-ion flow tube (ICP/SIFT) tandem mass spectrometer. The atomic cations were produced at approximately 5500 K in an ICP source and allowed to decay radiatively and to thermalize by collisions with argon and helium atoms prior to reaction. Rate coefficients and product distributions are reported for the reactions of fourth-row atomic cations from K(+) to Se(+), of fifth-row atomic cations from Rb(+) to Te(+) (excluding Tc(+)), of sixth-row atomic cations from Cs(+) to Bi(+), and of the lanthanide cations from La(+) to Lu(+) (excluding Pm(+)). Two primary reaction channels were observed: C-H bond insertion with elimination of H(2), and CH(4) addition. The bimolecular H(2) elimination was observed in the reactions of CH(4) with As(+), Nb(+), and some sixth-row metal cations, i.e., Ta(+), W(+), Os(+), Ir(+), Pt(+); secondary and higher-order H(2) elimination was observed exclusively for Ta(+), W(+), and Ir(+) ions. All other transition-metal cations except Mn(+) and Re(+) were observed to react with CH(4) exclusively by addition, and up to two methane molecules were observed to add sequentially to most transition-metal ions. CH(4) addition was also observed for Ge(+), Se(+), La(+), Ce(+), and Gd(+) ions, while the other main-group and lanthanide cations did not react measurably with methane.

14.
Anal Chim Acta ; 627(1): 148-53, 2008 Oct 03.
Article in English | MEDLINE | ID: mdl-18790138

ABSTRACT

Room-temperature reactions of the atomic cations Sr(+) and Rb(+) have been surveyed systematically with a variety of gases using an Inductively-Coupled Plasma/Selected-Ion Flow Tube (ICP/SIFT) tandem mass spectrometer. Rate coefficients and product distributions have been measured in He buffer gas at 0.35 Torr and 295 K for reactions of Sr(+) and Rb(+) with CH(3)F, CH(3)Cl, N(2)O, CO(2), CS(2), SF(6), D(2)O and NH(3). Rb(+) (s(0)) is seen to be quite inert with these molecules and reacts either slowly by molecule addition or not at all, while Sr(+) (s(1)) is much more reactive with all these 8 molecules, especially with CH(3)F, CH(3)Cl, N(2)O and SF(6). Sr(+) reacts with CH(3)F and SF(6) by F-atom transfer, with CH(3)Cl by Cl-atom transfer and with N(2)O by O-atom transfer, and the reaction rate coefficients are all quite high, k > or = 1.4x10(-11) cm(3) molecules(-1) s(-1). The extreme differences in reactivity with CH(3)F, SF(6), CH(3)Cl and N(2)O provide a chemical basis for the separation of isobaric interferences of (87)Rb(+) and (87)Sr(+) often encountered in ICP-MS. Among these four molecules, SF(6) exhibits the largest difference in reactivity, almost a factor of 10(4), and so is identified as the kinetically recommended reagent for the chemical resolution of the isobaric interference of (87)Rb(+) and (87)Sr(+).

15.
J Phys Chem A ; 111(35): 8561-73, 2007 Sep 06.
Article in English | MEDLINE | ID: mdl-17696503

ABSTRACT

Reactions of heavy water, D(2)O, have been measured with 46 atomic metal cations at room temperature in a helium bath gas at 0.35 Torr using an inductively coupled plasma/selected ion flow tube tandem mass spectrometer. The atomic cations were produced at ca. 5500 K in an ICP source and were allowed to decay radiatively and thermalize by collisions with Ar and He atoms prior to reaction. Rate coefficients and product distributions are reported for the reactions of fourth-row atomic cations from K+ to Se+, of fifth-row atomic cations from Rb+ to Te+ (excluding Tc+), and of sixth-row atomic cations from Cs+ to Bi+. Primary reaction channels were observed leading to O-atom transfer, OD transfer, and D2O addition. O-Atom transfer occurs almost exclusively (>or=90%) in the reactions with most early transition-metal cations (Sc+, Ti+, V+, Y+, Zr+, Nb+, Mo+, Hf+, Ta+, and W+) and to a minor extent (10%) with one main-group cation (As+). OD transfer is observed to occur only with three cations (Sr+, Ba+, and La+). Other cations, including most late transition and main-group cations, were observed to react with D2O exclusively and slowly by D2O addition or not at all. O-Atom transfer proceeds with rate coefficients in the range of 8.1 x 10(-13) (As+) to 9.5 x 10(-10) (Y+) cm3 molecule(-1)(s-1) and with efficiencies below 0.1 and even below 0.01 for the fourth-row atomic cations V+ (0.0032) and As+ (0.0036). These low efficiencies can be understood in terms of the change in spin required to proceed from the reactant to the product potential energy surfaces. Higher order reactions are also measured. The primary products, NbO+, TaO+, MoO+, and WO+, are observed to react further with D(2)O by O-atom transfer, and ZrO+ and HfO+ react further through OD group abstraction. Up to five D(2)O molecules were observed to add sequentially to selected M+ and MO+ as well as MO2+ cations and four to MO(2)D+. Equilibrium measurements for sequential D(2)O addition to M+ are also reported. The periodic variation in the efficiency (k/k(c)) of the first addition of D(2)O appears to be similar to the periodic variation in the standard free energy (DeltaG degrees) of hydration.

16.
Anal Chem ; 79(11): 4006-12, 2007 Jun 01.
Article in English | MEDLINE | ID: mdl-17487975

ABSTRACT

The kinetics and product distributions of the reactions of dimethyl disulfide (DMDS) have been investigated with a group of chemical background ions commonly observed in atmospheric pressure ionization (API) mass spectrometry (MS) in order to assess the value of this molecule in filtering (or "scrubbing") these ions by changing their mass/charge (m/z) ratio. The measurements were taken with a novel electrospray ionization/selected ion flow tube/QqQ tandem mass spectrometer. The background ions studied include those with m/z 42 (protonated acetonitrile, ACN), 83 (protonated ACN dimer), 99 (protonated phosphoric acid), 117 (water cluster of m/z 99), 131 (methanol cluster of m/z 99), 149 (protonated phthalic anhydride, formed from the phthalates), and 327 (protonated triphenyl phosphate). In addition, reactions of DMDS have been studied with two model analytes--protonated caffeine and doubly protonated bradykinin--in order to assess the selectivity of DMDS reactivity. All the measurements were taken at 295 +/- 2 K in helium buffer gas at a pressure of 0.35 +/- 0.01 Torr. DMDS was observed to react efficiently with m/z 42 (ACNH+), 149 (from phthalates), and 99 (protonated phosphoric acid), with k/kc=0.91, 0.47, and 0.38, respectively. Only proton transfer was observed with ACNH+, followed by the secondary reaction of [DMDSH]+ with DMDS to yield [CH3S-S(CH3)-SCH3]+. Ligation of DMDS was the dominant primary channel observed for the reaction of the m/z 149 background ion; however, some proton transfer also was observed. Both of these primary product ions react further with DMDS to yield [CH3S-S(CH3)-SCH3]+, the structure of which we have determined computationally using DFT calculations. Only the sequential ligation with two DMDS molecules was observed for the reaction of the m/z 99 ion. Reactions of DMDS with m/z 117 [H3PO4 + H + H2O]+ and m/z 131 [H3PO4 + H + MeOH]+ were observed to proceed with k/kc=0.71 and 0.058, respectively. Ligand substitution of DMDS for H2O predominated ( approximately 94%) over DMDS ligation ( approximately 6%) in the reaction with m/z 117, while only DMDS ligation was observed for the reaction of m/z 131 with DMDS. In contrast, the reactions of DMDS with ions of m/z 83 (protonated dimer of ACN) and 327 (protonated triphenyl phosphate) were extremely inefficient, with k/kc=0.0042 and 0.0079, respectively. The higher reactivity of DMDS toward ACNH+ (m/z 42) compared to (ACN)2H+ (m/z 83) is attributed to the lower proton affinity of the unsolvated ACN. The reactivity of DMDS toward the two model analyte ions studied-protonated caffeine and doubly protonated bradykinin-was negligible, with k/kc=0.0073 and 0.010, for the respective reactions. These results suggest that, under appropriate reagent pressure conditions, DMDS can be an appropriate reagent for chemically filtering out many common API-MS background ions, without significantly affecting the observed intensity of analyte peaks.

17.
Inorg Chem ; 45(24): 9646-53, 2006 Nov 27.
Article in English | MEDLINE | ID: mdl-17112259

ABSTRACT

Results are reported for experimental measurements of the room-temperature chemical reactions between ground-state Hg*+ ions and 16 important environmental and biological gases: SF6, CO, CO2, N2O, D2O, CH4, CH3F, O2, CH3Cl, OCS, CS2, NH3, C6F6, NO2, NO*, and C6H6. The inductively coupled plasma/selected-ion flow tube tandem mass spectrometer used for these measurements has provided both rate and equilibrium constants. Efficient electron transfer (>19%) is observed with CS2, NH3, C6F6, NO2, NO*, and C6H6, molecular addition occurs with D2O, CH4, CH3F, CH3Cl, and OCS, and SF6, CO, CO2, N2O, and O2 showed no measurable reactivity with Hg*+. Theory is used to explore the stabilities and structures of both the observed and unobserved molecular adducts of Hg*+, and reasonable agreement is obtained with experimental observations, given the uncertainties of the theory and experiments. A correlation is reported between the Hg*+ and proton affinities of the ligands investigated. Solvation of Hg*+ with formic acid was observed to increase the rate of electron transfer from NO* by more than 20%.

18.
J Phys Chem A ; 110(47): 12832-8, 2006 Nov 30.
Article in English | MEDLINE | ID: mdl-17125297

ABSTRACT

Gas-phase reactions of atomic lanthanide cations (excluding Pm+) have been surveyed systematically with CO2 and CS2 using an inductively coupled plasma/selected-ion flow tube (ICP/SIFT) tandem mass spectrometer. Observations are reported for reactions with La+, Ce+, Pr+, Nd+, Sm+, Eu+, Gd+, Tb+, Dy+, Ho+, Er+, Tm+, Yb+, and Lu+ at room temperature (295 +/- 2 K) in helium at a total pressure of 0.35 +/- 0.02 Torr. The observed primary reaction channels correspond to X-atom transfer (X = O, S) and CX2 addition. X-atom transfer is the predominant reaction channel with La+, Ce+, Pr+, Nd+, Gd+, Tb+, and Lu+, and CX2 addition occurs with the other lanthanide cations. Competition between these two channels is seen only in the reactions of CS2 with Nd+ and Lu+. Rate coefficient measurements indicate a periodicity in the reaction efficiencies of the early and late lanthanides. With CO2 the observed trends in reactivity across the row and with exothermicity follow trends in the energy required to achieve two unpaired non-f valence electrons by electron promotion within the Ln+ cation that suggest the presence of a kinetic barrier, in a manner much like those observed previously for reactions with isoelectronic N2O. In contrast, no such barrier is evident for S-atom transfer from the valence isolectronic CS2 molecule which proceeds at unit efficiency, and this is attributed to the much higher polarizability of CS2 compared to CO2 and N2O. Up to five CX2 molecules were observed to add sequentially to selected Ln+ and LnX+ cations.

19.
J Phys Chem A ; 110(36): 10607-18, 2006 Sep 14.
Article in English | MEDLINE | ID: mdl-16956243

ABSTRACT

Reactions of CH(3)F have been surveyed systematically at room temperature with 46 different atomic cations using an inductively coupled plasma/selected-ion flow tube tandem mass spectrometer. Rate coefficients and product distributions were measured for the reactions of fourth-period atomic ions from K(+) to Se(+), of fifth-period atomic ions from Rb(+) to Te(+) (excluding Tc(+)), and of sixth-period atomic ions from Cs(+) to Bi(+). Primary reaction channels were observed corresponding to F atom transfer, CH(3)F addition, HF elimination, and H(2) elimination. The early-transition-metal cations exhibit a much more active chemistry than the late-transition-metal cations, and there are periodic features in the chemical activity and reaction efficiency that maximize with Ti(+), As(+), Y(+), Hf(+), and Pt(+). F atom transfer appears to be thermodynamically controlled, although a periodic variation in efficiency is observed within the early-transition-metal cations which maximizes with Ti(+), Y(+), and Hf(+). Addition of CH(3)F was observed exclusively (>99%) with the late-fourth-period cations from Mn(+) to Ga(+), the fifth-period cations from Ru(+) to Te(+), and the sixth-period cations from Hg(+) to Bi(+) as well as Re(+). Periodic trends are observed in the effective bimolecular rate coefficient for CH(3)F addition, and these are consistent with expected trends in the electrostatic binding energies of the adduct ions and measured trends in the standard free energy of addition. HF elimination is the major reaction channel with As(+), while dehydrogenation dominates the reactions of W(+), Os(+), Ir(+), and Pt(+). Sequential F atom transfer is observed with the early-transition-metal cations, with the number of F atoms transferred increasing across the periodic table from two to four, maximizing at four for the group 5 cations Nb(+)(d(4)) and Ta(+)(d(3)s(1)), and stopping at two with V(+)(d(4)). Sequential CH(3)F addition was observed with many atomic cations and all of the metal mono- and multifluoride cations that were formed.


Subject(s)
Hydrocarbons, Fluorinated/chemistry , Metals/chemistry , Temperature , Cations/chemistry , Gases/chemistry , Hydrofluoric Acid/chemistry , Hydrogen/chemistry , Kinetics
20.
Chemphyschem ; 7(8): 1813-9, 2006 Aug 11.
Article in English | MEDLINE | ID: mdl-16810661

ABSTRACT

Reactions of atomic lanthanide cations (excluding Pm+) with D2O have been surveyed in the gas phase using an inductively coupled plasma/selected-ion flow tube (ICP/SIFT) tandem mass spectrometer to measure rate coefficients and product distributions in He at 0.35+/-0.01 Torr and 295+/-2 K. Primary reaction channels were observed corresponding to O-atom transfer, OD transfer and D2O addition. O-atom transfer is the predominant reaction channel and occurs exclusively with Ce+, Nd+, Sm+, Gd+, Tb+ and Lu+. OD transfer is observed exclusively with Yb+, and competes with O-atom transfer in the reactions with La+ and Pr+. Slow D2O addition is observed with early lanthanide cation Eu+ and the late lanthanide cations Dy+, Ho+, Er+ and Tm+. Higher-order sequential D2O addition of up to five D2O molecules is observed with LnO+ and LnOD+. A delay of more than 50 kcal mol(-1) is observed in the onset of efficient exothermic O-atom transfer, which suggests the presence of kinetic barriers of perhaps this magnitude in the exothermic O-atom transfer reactions of Dy+, Ho+, Er) and Tm+ with D2O. The reaction efficiency for O-atom transfer is seen to decrease as the energy required to promote an electron to make two non-f electrons available for bonding increases. The periodic trend in reaction efficiency along the lanthanide series matches the periodic trend in the electron-promotion energy required to achieve a d1s1 or d2 excited electronic configuration in the lanthanide cation, and also the periodic trends across the lanthanide row reported previously for several alcohols and phenol. An Arrhenius-like correlation is also observed for the dependence of D2O reactivity on promotion energy for early lanthanide cations, and exhibits a characteristic temperature of 2600 K.

SELECTION OF CITATIONS
SEARCH DETAIL
...